2.1. Characterization of VR and VREP
The synthetic route employed to obtain the VREP via the eco-friendly L-proline-catalyzed approach is outlined in Scheme S1. Following the preparation of the solid VR and VREP samples, their chemical structures were confirmed by 1H nuclear magnetic resonance (NMR), 13C NMR (Fig. S1 and S2), and Fourier transform infrared (Ft-ir) spectroscopy (Fig. S3). As observed from the 1H NMR spectra of VR and VREP (Fig. S1a and S2a), doublets at 6.57 and 7.45 ppm confirmed the presence of alpha and beta protons of the unsaturated ketone moiety, while the peaks at 6.75, 6.91, 7.01, 7.06, and 7.09 ppm corresponded to protons of the aromatic benzene ring. However, upon comparison of the 1H NMR spectra of VR and VREP, the peaks at 4.54 and 5.88 ppm corresponding to the phenolic protons of VR disappeared, while eight new peaks derived from the epoxy group of VPEP appeared at 2.74, 2.89, 3.25, 3.38, 3.92, 4.04, 4.18, and 4.30 ppm, thereby confirming that the reaction was successful. In addition, the chemical shifts and carbon numbers in 13C NMR spectra matched the carbons of VR and VREP, respectively (Fig, S1b and S2b). Furthermore, from the Ft-ir results, peaks corresponding to unsaturated ketone moieties (i.e., at 1650 and 1620 cm− 1) were confirmed for both compounds. As expected, the VR hydroxyl peak at 3300–3200 cm− 1 disappeared, while an oxirane peak (910 cm− 1) and ether peaks (1140 and 1210 cm− 1) corresponding to the epoxy group appeared in the spectrum of VREP, thus confirming the successfully coupling reaction.
2.2 Curing behavior analysis of the mixed VREP/DDS system
The reaction between VREP and DDS is outlined in Scheme S2. To determine the gelling time required to obtain the crosslink network, the complex viscosity and shear modulus of the VREP and DDS mixture were measured while increasing the temperature of the system at a rate of 2 ℃ min− 1. As shown in Fig. 2a, the complex viscosity tended to remain constant until the temperature increased rapidly at approximately 33 min (at 200 ℃) due to curing. Subsequently, the viscosity increased at a slower rate as the viscosity inflection point was reached. Although the gelling time was close to this inflection point, it was difficult to precisely determine the exact gelling time. Thus, the shear storage and loss moduli (G' and G'') was measured under the same conditions to obtain an accurate gelling time. In general, the crossover point where the G' and G'' values intersect is defined as the gel time [18]. Thus, as shown in Fig. 2b, the crossover point was located at 33 min, which is consistent with the inflection point of the complex viscosity–time graph, and therefore reflects the gelling time of the VREP and DDS mixture.
To obtain a deeper understanding of this system, we investigated the curing behavior of the VREP and DDS mixture by differential scanning calorimetry (DSC). For comparison, DSC of VREP alone was also carried out (Fig. S4), and an endothermic peak was observed at 142.8 ℃ during the first heating, which corresponded to the melting point (Tm) of VREP. Figure 2c shows the DSC plot of the VREP and DDS mixture, wherein a slightly lower Tm of 131 ℃ was recorded. In addition, an exothermic peak was observed between 175 and 245 ℃, which represented the curing temperature range of the mixture. After the first heating, no endothermic or exothermic peaks were observed, indicating that complete curing had been achieved during the first cooling process. Thus, to explore the optimal curing temperature conditions in more detail, the enthalpy was calculated at each temperature point by dividing the temperature range into five equal stages between the starting onset point (175 ℃) and the maximum peak (215 ℃) in isothermal mode over 1 h (Fig. 2d). It was found that the highest enthalpy (305 J g− 1) correlated to a temperature of 185 ℃, and no heat flow was detected after 1 h. These results indicate that a gelling time of 33 min and a curing temperature and time of 185 ℃ and 1 h, respectively, were optimal for the VREP and DDS mixture.
2.3 Mechanical and thermal properties of the cured epoxy resin based on VREP and DDS
Based on the above curing behavior analysis result, the mixture of VREP and DDS was pre-cured for 1 h at the onset point temperature (175 ℃) prior to post-curing for 3 h at the optimal temperature (185 ℃) to ensure complete curing. The specimens are shown in the insets of Fig. 3a and 3b, and the tensile and flexural stress–strain curves of the cured epoxy resin based on VREP and DDS (abbreviated as VREP/DDS) are shown in Fig. 3a and 3b. From these plots, the tensile strength and Young’s modulus of VREP/DDS were determined to be 58 MPa and 2.9 GPa, respectively. The tensile strain gauge was recorded at 2.7%. Thus, our system exhibited a tensile strength similar to that of thermosetting plastics based on BPA-type epoxy resins (i.e., the diglycidyl ether of bisphenol A epoxy resin combined with DDS, DGEBA/DDS), with tensile strengths in the range of 57.2–59.4 MPa, in addition to a greater Young’s modulus (c.f., 0.56 GPa for DGEBA/DDS) [19]. Furthermore, the flexural strength and flexural modulus of VREP/DDS were 183 MPa and 3.6 GPa, respectively, and the flexural strain gauge was recorded at 4.8%. These results indicate a high flexural strength and flexural modulus compared with the DGEBA/DDS system (i.e., 79.5 MPa and 1.03 GPa, respectively), [19] thereby confirming the slightly superior mechanical strength of VREP/DDS compared to DGEBA/DDS.
Dynamic mechanical analysis (DMA) was then employed to evaluate the storage modulus, loss modulus, and tanδ. As shown in Fig. 3c, the storage modulus was 5351 MPa and the loss modulus was 432 MPa at the maximum peak located at 140 ℃. Based on these data, it can be observed that the maximum peak from the calculated tanδ graph occurred at 151 ℃, and this value corresponded to the glass transition temperature (Tg). Furthermore, the middle flection point of the heat flow, which also indicates Tg, was detected at 148 ℃ in the DSC plot of VREP/DDS (Fig. 3d), thereby confirming similar Tg values obtained by both DMA and DSC analysis [19].
Furthermore, we investigated the crosslink density of the VREP/DDS system based on these DMA and DSC analysis data, as outlined in Eq. (1):
\(\rho =\frac{{E}^{{\prime }}}{3RT} \left(1\right)\)
where ρ, E', T and R are the crosslink density (mol cm− 3), tensile storage modulus (J cm− 3) at the rubbery plateau region, gas constant (8.314772 J K− 1 mol− 1), and temperature (K) at determined E', respectively. The value of E' was selected based on the point at which it became constant at 191 ℃; this temperature is higher 40 degree Celsius than Tg [20], thereby giving an E' value of 157.9 J cm− 3 at 191 ℃. Thus, the crosslink density of the VREP/DDS system was determined to be 0.0136 mol cm− 3 based on the above equation. Since the crosslink density of DGEBA/DDS has been reported to be ~ 0.0118 mol cm− 3 [20], it is clear that the crosslink density of our VREP/DDS system is higher, exhibiting a higher storage modulus and superior mechanical properties compared to those of DGEBA/DDS. Thus, it can be inferred that the proposed VREP/DDS system is a potential replacement to the BPA-type epoxy.
2.4 Application of the VREP/DDS system to CFRP
Subsequently, to evaluate the potential application of the VREP/DDS system to CFRP, the interfacial shear strength (IFSS) was measured using the pull-out test method to determine the resistance to shear force between the cured epoxy resin and carbon fiber (CF). Figure 4a shows a photograph of the pull-out test setup and the obtained IFSS for the VREP/DDS and DGEBA/DDS systems, wherein values of 35.0 and 40.8 MPa were obtained, respectively. Although the IFSS of the VREP/DDS system was slightly lower than that of the DGEBA/DDS system, this was considered acceptable as the IFSS between BPA-type epoxy resins and CF typically ranges from 30 to 40 MPa [21].
During the fabrication of CFRP, it is necessary to evaluate how well the resin is impregnated into the CF, since a low degree of impregnation can create pores and voids, thereby resulting in inferior mechanical properties. It is therefore advantageous to impregnate into CF at a low complex viscosity; since VREP is a solid epoxy compound that cannot be impregnated at 25 ℃, its impregnation must be carried out in the molten state, at which point the viscosity should be evaluated. Figure 4b shows a plot of the complex viscosity against temperature. As indicated, at a temperature of ≤ 170 ℃, the complex viscosity was maintained below 110 mPa∙s, although it increased rapidly beyond 170 ℃. Thus, to reduce void formation, impregnation was carried out at 170 ℃, i.e., before the sharp increase in the complex viscosity. The CFRP was prepared using the VREP and DDS mixture and a CF weave fabric (VREP/DDS-CFRP) as described previously under relatively low complex viscosity conditions for dwelling (i.e., 170 ℃, 30 min.) and with curing at 185 ℃ for 4 h. Specimens for the tensile and ILSS tests were fabricated by laminating 3 and 10 piles of the CF fabric, respectively. Figure 4c shows photographs of the fabrication process and the prepared VREP/DDS-CFRP sample.
Figure 4d showed a scanning electron microscope image of how resin was impregnated in the CF. From the image, it can be seen that it was impregnated well without any empty space. Subsequently, to evaluate the number of voids, the cut cross-section (i) and the top surface (ii) were analyzed by X-ray microscopy (XRM). Figure 4e showed computed tomography (CT) images for x and y axis directions, and a 3D rendering reconstructed from the scanned CT images. Small voids marked with red arrows could be identified in the CT image (see Fig. 4e-i, ii), and we calculated the void volume fraction by summing the volumes of all voids present in the VREP/DDS-CFRP from the 3D rendering. Thus, a fiber volume fraction of 60% was obtained in addition to a void volume fraction of 0.36%. Importantly, this value is within the limit of ≤ 1% specified for CFRP materials exhibiting high mechanical properties [22]. Therefore, although VREP is a solid epoxy resin that cannot easily to impregnate CF, our results indicated that no significant reduction in the mechanical properties took place following melting and impregnating.
The mechanical properties of the VREP/DDS-CFRP system were then evaluated as shown in Fig. 4f and 4g, and the tensile strength and ILSS results were obtained. The corresponding specimens are shown in the insets, and the obtained results are listed in Table S1. More specifically, the average tensile strength and Young’s modulus were 957 MPa and 77 GPa, respectively, and the average ILSS was 49 MPa. These results are similar to those of the CFRP prepared using the DGEBA-type epoxy resin [23], which indicates the potential of VREP for application in CFRP products to replace BPA-type epoxy resins.
2.5 Chemical degradability of the cured VREP/DDS-containing epoxy resin
As described above, the molecular structure was designed considering the retro-aldol condensation reaction, which initially involves conversion of the alkene group of an unsaturated ketone into a secondary alcohol. The resulting compound then reacts under heating and basic conditions in solvents such as sodium hydroxide (NaOH) solution, which promote retro-aldol reaction [24, 25]. Thus, we oxidized VREP/DDS using hydrochloric acid (HCl) solution and selected sodium hypochlorite (NaOCl) solution as the basic solvent due to its ability to generate NaOH at room temperature over a pH range of 12–15 [26].
Figure 5a shows the decomposition process and the resulting visible changes that occurred in the VREP/DDS sample. More specifically, following the oxidation process (step 1) that was carried out in a refluxing 1.0 M HCl solution (100 ℃) over 12 h, the sample color was changed from yellow to brown. The brown VREP/DDS sample was then reacted in a 1.0 M NaOCl solution at 100 ℃ for 24 h (step 2), during which time the solution color changed to yellow, and the complete dissolution and decomposition of VREP/DDS was observed. To obtain a deeper understanding of the mechanism involved in this process, FT-IR spectroscopy was carried out (Fig. 5b). Following the oxidation process, the peak corresponding to the alkene group of the unsaturated ketone moiety (C = C, 1650 cm− 1) disappeared, and a new secondary alcohol peak was observed (C‒O, 1168 cm− 1). In addition, after step 2, the secondary alcohol peak disappeared and new aldehyde peaks were observed (C–H, 2740 and 2789 cm− 1) [27].
Based on these FT-IR results, we propose a mechanism for the chemical degradation process of VREP/DDS (Fig. 5c). More specifically, in step 1, a proton (H+) generated by HCl attacks the alpha site of the unsaturated ketone alkene moiety. Subsequently, a water molecule attacks the cation and H+ is released to form a secondary alcohol. In step 2, an OH‾ ion generated from the NaOCl solution abstracts a proton from a secondary alcohol to form water, and the retro-aldol reaction is promoted by heating to decompose the structure of VREP/DDS into an aldehyde and a ketone, thereby confirming the chemical degradability of this system.
2.6 Chemical recycling of VREP/DDS-CFRP
Following the confirmation of the chemical degradability of VREP, we applied the same procedure to investigate the decomposition of VREP/DDS-CFRP, as outlined in Fig. 6a. As shown, in step 1, the color of the VREP/DDS-CFRP sample changed to brown caused by oxidation, while a yellow solution formed following step 2. After washing of the solid with water and subsequent drying, the reclaimed CF (r-CF) was obtained.
From the viewpoint of decomposition kinetics, the degrees of decomposition were investigated at 1 h intervals for VREP/DDS-CFRP and for the DGEBA/DDS-CFRP sample prepared according to the same procedure. Based on the obtained data, the degree of decomposition was calculated in each case, and the decomposition rate constant was determined as follows:
In Eq. (2), [NaOCl] and [H2O] can be considered constants because excess amounts of H2O and NaOCl were used in the decomposition process. As a result, Eq. (3) is obtained, wherein [EP] corresponds to a weight (g) rather than a concentration due to the fact that the EP specimen was a solid [28]. All results are presented in Table S2. Figure 6b and S5a show the degrees of decomposition of VREP/DDS-CFRP and DGEBA/DDS-CFRP, respectively, over time, wherein it can be seen that the decomposition degree of DGEBA/DDS-CFRP remained at almost 0% after 4 h, while 100% decomposition was observed after 4 h for VREP/DDS-CFRP. In addition, Fig. 6c and S5b were plotted to obtain the reaction rate constants (k) for VREP/DDS-CFRP and DGEBA/DDS-CFRP, which were determined to be 1.4 × 10− 2 and 0.064 × 10− 2 min− 1, respectively; it was confirmed that the decomposition reactions of both specimens were of the first order with respect to the residual EP concentration. As expected, a rate constant of essentially zero was obtained for DGEBA/DDS-CFRP, while VREP/DDS-CFRP gave a rate constant ~ 20 times higher, thereby confirming the superior chemical degradability of the latter due to the retro-aldol reaction of VREP.
Finally, the insets of Fig. 6d and S6 show the FE-SEM images of the r-CF sample at 500× and 1000× magnification, respectively. From these images, it was observed that no epoxy resin remained on the r-CF surface, thereby confirming its successful decomposition. The tensile strength of the r-CF sample was then measured to determine whether mechanical properties of the r-CF were retained compared to those of the v-CF. More specifically, the tensile strength and Young’s modulus of the r-CF sample were 4.2 and 226.1 GPa, respectively, which represent at 13% loss compared to the corresponding values of the v-CF sample, i.e., 4.8 and 260.4 GPa, respectively (Fig. 6d), thereby confirming a good degree of mechanical performance retention following the decomposition process.